\documentclass[reqno]{amsart} \usepackage{hyperref} \AtBeginDocument{{\noindent\small 2006 International Conference in Honor of Jacqueline Fleckinger. \newline \emph{Electronic Journal of Differential Equations}, Conference 16, 2007, pp. 35--58. \newline ISSN: 1072-6691. URL: http://ejde.math.txstate.edu or http://ejde.math.unt.edu \newline ftp ejde.math.txstate.edu (login: ftp)} \thanks{\copyright 2007 Texas State University - San Marcos.} \vspace{9mm}} \begin{document} \setcounter{page}{35} \title[\hfilneg EJDE/Conf/16 \hfil Ground-state compactness] {Compactness for a Schr\"odinger operator\\ in the ground-state space over $\mathbb{R}^N$} \author[B. Alziary, P. Tak\'a\v{c} \hfil EJDE/Conf/16 \hfilneg] {B\'en\'edicte Alziary, Peter Tak\'a\v{c}} % in alphabetical order \address{B\'en\'edicte Alziary \newline CEREMATH -- UMR MIP, Universit\'e Toulouse~1 (Sciences Sociales), 21 All\'ees de Brienne, F--31000 Toulouse Cedex, France} \email{alziary@univ-tlse1.fr} \address{Peter Tak\'a\v{c} \newline Institut f\"ur Mathematik, Universit\"at Rostock, Universit\"atsplatz 1, D--18055 Rostock, Germany} \email{peter.takac@uni-rostock.de} \thanks{Published May 15, 2007.} \subjclass[2000]{47A10, 35J10, 35P15, 81Q15} \keywords{ Ground-state space; compact resolvent; Schr\"odinger operator; \hfill\break\indent monotone radial potential; maximum and anti-maximum principle; \hfill\break\indent comparison of ground states; asymptotic equivalence} \dedicatory{Dedicated to Jacqueline Fleckinger on the occasion of \\ an international conference in her honor} \begin{abstract} We investigate the compactness of the resolvent $(\mathcal{A} - \lambda I)^{-1}$ of the Schr\"odinger operator $\mathcal{A} = - \Delta + q(x)\bullet$ acting on the Banach space $X$, \begin{equation*} X = \{ f\in L^2(\mathbb{R}^N): f / \varphi\in L^\infty(\mathbb{R}^N) \} ,\quad \| f\|_X = \mathop{\rm ess\,sup}_{\mathbb{R}^N} (|f| / \varphi)\, , \end{equation*} $X\hookrightarrow L^2(\mathbb{R}^N)$, where $\varphi$ denotes the ground state for $\mathcal{A}$ acting on $L^2(\mathbb{R}^N)$. The potential $q: \mathbb{R}^N\to [q_0,\infty)$, bounded from below, is a ``relatively small'' perturbation of a radially symmetric potential which is assumed to be monotone increasing (in the radial variable) and growing somewhat faster than $|x|^2$ as $|x|\to \infty$. If $\Lambda$ is the ground state energy for $\mathcal{A}$, i.e.\ $\mathcal{A}\varphi = \Lambda\varphi$, we show that the operator $(\mathcal{A} - \lambda I)^{-1} : X\to X$ is not only bounded, but also compact for $\lambda\in (-\infty, \Lambda)$. In particular, the spectra of $\mathcal{A}$ in $L^2(\mathbb{R}^N)$ and $X$ coincide; each eigenfunction of $\mathcal{A}$ belongs to $X$, i.e., its absolute value is bounded by $\mathrm{const}\cdot \varphi$. \end{abstract} \maketitle \numberwithin{equation}{section} \newtheorem{theorem}{Theorem}[section] \newtheorem{corollary}[theorem]{Corollary} \newtheorem{lemma}[theorem]{Lemma} \newtheorem{proposition}[theorem]{Proposition} \newtheorem{remark}[theorem]{Remark} \newtheorem{example}[theorem]{Example} \section{Introduction} \label{s:Intro} We investigate the compactness of the resolvent $(\mathcal{A} - \lambda I)^{-1}$ of the Schr\"odinger operator % \begin{equation} \mathcal{A}\equiv \mathcal{A}_q\stackrel{{\rm def}}{=} - \Delta + q(x)\bullet \label{op:Schroed_lam} \end{equation} % acting not only on the standard Hilbert space $L^2(\mathbb{R}^N)$, but also on the Banach space $X$, % \begin{equation} X\equiv X_q\stackrel{{\rm def}}{=} \{ f\in L^2(\mathbb{R}^N): f / \varphi\in L^\infty(\mathbb{R}^N) \} , \label{def:X} \end{equation} % or on its predual space $X^\odot = L^1(\mathbb{R}^N; \varphi \,\mathrm{d}x)$. Here, $\varphi\equiv \varphi_q$ denotes the (normalized) ground state of $\mathcal{A}$. The electric potential $q(x)$ is assumed to be a continuous function $q: \mathbb{R}^N\to \mathbb{R}$ such that % \begin{equation} q_0\stackrel{{\rm def}}{=} \inf_{\mathbb{R}^N} q > 0 \quad\mbox{and }\quad q(x)\to +\infty \mbox{ as } |x|\to \infty . \label{cond:q_0} \end{equation} % We denote by $\Lambda\equiv \Lambda_q$ the principal eigenvalue of the operator $\mathcal{A}$ (also called the ground state energy); hence, $\mathcal{A}\varphi = \Lambda\varphi$. It follows from \eqref{cond:q_0} that $(\mathcal{A} - \lambda I)^{-1}$ is a compact linear operator acting on $L^2(\mathbb{R}^N)$ for every $\lambda\in (-\infty, \Lambda)$. Furthermore, $(\mathcal{A} - \lambda I)^{-1}$ is bounded as an operator acting on $X$, by a standard argument using the weak maximum principle. This operator, in general, is {\it not } compact on $X$; for instance, not for $q(x) = |x|^2$ ($|x|\geq r_0 > 0$). As a direct consequence of the Riesz\--Schauder theory for compact linear operators ({Edwards} \cite[{\S}9.10, pp.\ 677--682]{Edwards} or {Yosida} \cite[Chapt.~X, Sect.~5, pp.\ 283--286]{Yosida}), if $(\mathcal{A} - \lambda I)^{-1}$ happens to be compact on $X$ then, for instance, all $L^2(\mathbb{R}^N)$\--eigenfunctions of $\mathcal{A}$ actually belong to $X$. Moreover, one can prove an anti\--maximum principle as well. We refer to {Alziary}, {Fleckinger}, and {Tak\'a\v{c}} \cite[Theorem 2.1, p.~128]{AFT-1} for $N=2$, \cite[Theorem 2.1, p.~365]{AFT-2} for $N\geq 2$, and to \cite[Sect.~3]{AFT-3} for further details in applications of the compactness of $(\mathcal{A} - \lambda I)^{-1}$ on~$X$. The corresponding results, under different hypotheses, are stated in Section~\ref{s:Main} below. In the work reported in the present article we persue the search (started in {Alziary}, {Fleckinger}, and {Tak\'a\v{c}} \cite{AFT-3}) for finding ``reasonable'' {\it sufficient conditions } on the potential $q(x)$ that guarantee the compactness of $(\mathcal{A} - \lambda I)^{-1}$ on $X$. Different sufficient conditions on $q$, which also force $(\mathcal{A} - \lambda I)^{-1}$ compact on $X$, are formulated in \cite[Theorem 3.2(a)]{AFT-3}. We will take advantage of some of the recent results obtained in \cite{AFT-3} to treat potentials $q$ satisfying % \begin{equation} \label{ineq:Q_1(r) 0$ -- a constant), then $v / \varphi\in L^{\infty}(\mathbb{R}^N)$ holds for {\it every } eigenfunction $v$ of $\mathcal{A}$, again by results from {Davies} \cite{Davies}, Corollary 4.5.5 (p.~122) combined with Lemma 4.2.2 (p.~110) and Theorem 4.2.3 (p.~111). We refer to {Davies} and {Simon} \cite[Theorem 6.3, p.~359]{DavSimon} and {M.~Hoffmann\--Ostenhof} \cite[Theorem 1.4(i), p.~67]{Hoffmann} for the same result under much weaker restrictions on $q(x)$. In our present article we impose similar restrictions. Now assume $\lambda\in \mathbb{C}$, $f\in L^2(\mathbb{R}^N)$, $f\geq 0$ and $f\not\equiv 0$ in~$\mathbb{R}^N$, and let $u\in L^2(\mathbb{R}^N)$ be a solution of the equation % \begin{equation} - \Delta u + q(x) u = \lambda u + f(x) \quad\mbox{in }\, L^2(\mathbb{R}^N) \label{e:Schroed_lam} \end{equation} % in the sense of distributions on $\mathbb{R}^N$. A related sufficient condition on $q$ is imposed in {Alziary} and {Tak\'a\v{c}} \cite[Theorem 2.1, p.~284]{AlziaryTak} to obtain % \begin{equation} u\geq c\varphi\quad\mbox{in $\mathbb{R}^N$ ({\it $\varphi$-positivity }) } \label{ineq:phi_+} \end{equation} % for $\lambda < \Lambda$, with some constant $c>0$. Somewhat stronger sufficient conditions on $q$ and $f$ guarantee also % \begin{equation} u\leq - c\varphi\quad\mbox{in $\mathbb{R}^N$ ({\it $\varphi$-negativity }) } \label{ineq:phi_-} \end{equation} % provided $\Lambda < \lambda < \Lambda + \delta$, where $\delta\equiv \delta(f) > 0$ is sufficiently small, $c>0$; see {Alziary}, {Fleckinger}, and {Tak\'a\v{c}} \cite[Theorem 2.1, p.~128]{AFT-1} for $N=2$ and \cite[Theorem 2.1, p.~365]{AFT-2} for $N\geq 2$. Moreover, in both inequalities \eqref{ineq:phi_+} and \eqref{ineq:phi_-} we have $c\equiv c(\lambda)\to +\infty$ as $\lambda\to \Lambda$. Again, the harmonic oscillator $q(x)\equiv |x|^2$ and a suitably chosen positive function $f$ provide easy counterexamples to both, \eqref{ineq:phi_+} and \eqref{ineq:phi_-}. From the proof of Theorem \ref{thm-X-compact}, Part~(a), we will derive \eqref{ineq:phi_+} whenever $f\in X^\odot$, $0\leq f\not\equiv 0$ in $\mathbb{R}^N$, and $\lambda < \Lambda$. This result is stated as Theorem \ref{thm-Positive-q(x)}. Directly from Theorem \ref{thm-X-compact}, Part~(c), we will derive also \eqref{ineq:phi_-} whenever $f\in X$, $\int_{\mathbb{R}^N} f\varphi \,{\rm d}x > 0$, and $\Lambda < \lambda < \Lambda + \delta$ ($\delta > 0$ -- small enough). This is the anti\--maximum principle in Theorem \ref{thm-Antimax-q(x)}. The proof of Theorem \ref{thm-X-compact} is based on the asymptotic equivalence of $\varphi_{Q_1}(|x|)$ and $\varphi_{Q_2}(|x|)$ as $|x|\to \infty$, for $Q_1$ and $Q_2$ as described above. An important ingredient here is Lemma \ref{lem-Titchmarsh} which generalizes {Titchmarsh}' lemma \cite[Sect.~8.2, p.~165]{Titchmarsh} applied in {Alziary} and {Tak\'a\v{c}} \cite[Lemma 3.2, p.~286]{AlziaryTak}, and in {Alziary}, {Fleckinger}, and {Tak\'a\v{c}} %\cite[Lemma 3.2, p.~132]{AFT-1} and \cite[Lemma 3.2, p.~366]{AFT-2} \cite[p.~132]{AFT-1} and \cite[p.~366]{AFT-2}, with a slightly different class of potentials $Q_1(r)$ and $Q_2(r)$. In {Alziary}, {Fleckinger}, and {Tak\'a\v{c}} \cite[Lemma 4.1]{AFT-3} the authors use a WKB\--type asymptotic formula due to {Hartman} and {Wintner} \cite[eq.\ (xxv), p.~49]{Hartman-Wint}. Asymptotic estimates for radially symmetric solutions of the Schr\"odinger equation with $q(x) = Q_j(|x|)$ ($j=1,2$) are combined with standard comparison results for solutions with different, but pointwise ordered (nonradial) potentials in order to control the asymptotic behavior of these solutions at infinity, and thus retain the compactness of the resolvent from the radially symmetric case ({\rm Proposition~\ref{prop-q_1 0 \quad\mbox{and}\quad q(x)\to +\infty \quad \mbox{as } |x|\to \infty . %\label{cond:q_0} \end{equation*} % We interpret equation \eqref{e:Schroed_lam} as the operator equation $\mathcal{A} u = \lambda u + f$ in $L^2(\mathbb{R}^N)$, where the Schr\"odinger operator \eqref{op:Schroed_lam}, % \begin{equation*} \mathcal{A}\equiv \mathcal{A}_q\stackrel{{\rm def}}{=} - \Delta + q(x)\bullet \quad\mbox{on }\, L^2(\mathbb{R}^N) , %\label{op:Schroed_lam} \end{equation*} % is defined formally as follows: We first define the quadratic (Hermitian) form % \begin{equation} \mathcal{Q}_q(v,w)\stackrel{{\rm def}}{=} \int_{\mathbb{R}^N} \left( \nabla v\cdot \nabla\bar{w} + q(x) v\bar{w} \right)\,{\rm d}x \label{e:q-form} \end{equation} % for every pair $v,w\in \mathcal{V}_q$ where % \begin{equation} \mathcal{V}_q\stackrel{{\rm def}}{=} \{ f\in L^2(\mathbb{R}^N): \mathcal{Q}_q(f,f) < \infty \} . \label{e:q-space} \end{equation} % Then $\mathcal{A}$ is defined to be the Friedrichs representation of the quadratic form $\mathcal{Q}_q$ in $L^2(\Omega)$; $L^2(\Omega)$ is endowed with the natural inner product % \[ (v,w)_{ L^2(\mathbb{R}^N) }\stackrel{{\rm def}}{=} \int_{\mathbb{R}^N} v\bar{w} \,{\rm d}x ,\quad v,w\in L^2(\Omega) . \] % This means that $\mathcal{A}$ is a positive definite, selfadjoint linear operator on $L^2(\Omega)$ with domain $\mathop{\rm dom}(\mathcal{A})$ dense in $\mathcal{V}_q$ and % \[ \int_{\mathbb{R}^N} (\mathcal{A} v)\bar{w} \,{\rm d}x = \mathcal{Q}_q(v,w) \quad\mbox{for all } v,w\in \mathop{\rm dom}(\mathcal{A}) ; \] % see {Kato} \cite[Theorem VI.2.1, p.~322]{Kato}. Notice that $\mathcal{V}_q$ is a Hilbert space with the inner product $(v,w)_q = \mathcal{Q}_q(v,w)$ and the norm $\| v\|_{\mathcal{V}_q} = ((v,v)_q)^{1/2}$. The embedding $\mathcal{V}_q\hookrightarrow L^2(\mathbb{R}^N)$ is compact, by \eqref{cond:q_0}. The principal eigenvalue $\Lambda\equiv \Lambda_q$ of the operator $\mathcal{A}\equiv \mathcal{A}_q$ can be obtained from the Rayleigh quotient % \begin{equation} \Lambda\equiv \Lambda_q = \inf \left\{ \mathcal{Q}_q(f,f) : f\in \mathcal{V}_q\ \mbox{ with }\ \| f\|_{L^2(\mathbb{R}^N)} = 1 \right\} , \quad \Lambda > 0 . \label{e:Lambda} \end{equation} % This eigenvalue is simple with the associated eigenfunction $\varphi\equiv \varphi_q$ normalized by $\varphi > 0$ throughout $\mathbb{R}^N$ and $\|\varphi\|_{L^2(\mathbb{R}^N)} = 1$; $\varphi$ is a minimizer for the Rayleigh quotient above. The reader is referred to {Edmunds} and {Evans} \cite{EdmEvans} or {Reed} and {Simon} \cite[Chapt.\ XIII]{RSimon-IV} for these and other basic facts about Schr\"odinger operators. We set $r = |x|$ for $x\in \mathbb{R}^N$, so $r\in \mathbb{R}_+$, where $\mathbb{R}_+\stackrel{{\rm def}}{=} [0,\infty)$. If $q$ is a radially symmetric potential, $q(x) = q(r)$ for $x\in \mathbb{R}^N$, then also the eigenfunction $\varphi$ must be radially symmetric. This follows directly from $\Lambda$ being a simple eigenvalue. Since our technique is based on a perturbation argument for a relatively small perturbation of a radially symmetric potential, which is assumed to satisfy certain differentiability and growth conditions in the radial variable $r = |x|$, $r\in \mathbb{R}_+$, we bound the potential $q: \mathbb{R}^N\to \mathbb{R}$ by such radially symmetric potentials from below and above. In order to formulate our hypotheses on the potential $q(x)$, $x\in \mathbb{R}^N$, we first introduce the following class \eqref{class:Q(r)} of {\it auxiliary functions } $Q(r)$ of $r=|x|\geq 0$: \begin{enumerate} % \item[(Q)] \makeatletter \def\@currentlabel{Q}\label{class:Q(r)} \makeatother % $Q: \mathbb{R}_+\to (0,\infty)$ is a continuous function that is monotone increasing in some interval $[r_0,\infty)$, $0 < r_0 < \infty$, and satisfies % \begin{equation} \int_{r_0}^\infty Q(r)^{-1/2} \,\mathrm{d}r < \infty . \label{e:int-Q^-1/2} \end{equation} % \end{enumerate} In particular, we have $Q'(r)\geq 0$ for a.e.\ $r\geq r_0$, and $Q(r)\to \infty$ as $r\to \infty$. Example \ref{exam-q(r):mono} in the Appendix, {\S}\ref{ss:Examples}, is essential for understanding potentials of class \eqref{class:Q(r)}. The potential $q(x) = q(|x|)$ exhibited in this example belongs to class \eqref{class:Q(r)}, but does {\it not } belong to the analogue of this class defined in {Alziary}, {Fleckinger}, and {Tak\'a\v{c}} \cite{AFT-3}. There, instead of $Q$ monotone increasing in $[r_0,\infty)$, it is required that there be a constant $\gamma$, $1 < \gamma\leq 2$, such that % \[ \int_{r_0}^\infty \Big| \frac{{\rm d}}{{\rm d}r} \Big( Q(r)^{-1/2} \Big) \Big|^{\gamma} Q(r)^{1/2} \,{\rm d}r < \infty . \label{e:int-deriv} \] % As a consequence, in our present paper we do not need to employ the Hartman\--Wintner asymptotic formula from \cite{Hartman-Wint}, eq.\ (xxv) on p.~49 or eq.\ (158) on p.~80. Condition \eqref{e:int-deriv} originally appeared in the work of {Hartman} and {Wintner} \cite{Hartman-Wint}, on p.~49, eq.\ (xxiv), and on p.~80, eq.\ (157). We impose the following hypothesis on the growth of $q$ from below: \begin{enumerate} % \item[(H1)] \makeatletter \def\@currentlabel{H1}\label{hyp:growth:q(r)} \makeatother % There exists a function $Q: \mathbb{R}_+\to (0,\infty)$ of class {\rm (Q)} such that % \begin{equation} q(x) - \Lambda + \frac{(N-1)(N-3)}{4 r^2} \geq Q(r) > 0 \quad\mbox{holds for all } |x| = r > r_0 . \label{growth:q(r)} \end{equation} % \end{enumerate} \begin{remark}\label{rem-growth:q(r)} \rm The term $- \Lambda + \frac{(N-1)(N-3)}{4 r^2}$ has been added for convenience only; it may be left out by replacing $Q(r)$ by $Q(r) + \Lambda + 1$ if $N\geq 1$ and taking also $r_0 > 0$ large enough if $N=2$. \end{remark} In several results we need stronger hypotheses than \eqref{hyp:growth:q(r)}. Writing $x = r x'$ ($x\in \mathbb{R}^N\setminus \{\mathbf{0}\}$) with the radial and azimuthal variables $r = |x|$ and $x'= x/|x|$, respectively, we frequently impose the following stronger restrictions on the growth of $q(x)$ in $r$ and the variation of $q(x)$ in $x'$: We assume that % \begin{enumerate} % \item[(H2)] \makeatletter \def\@currentlabel{H2}\label{hyp:growth:q(x)} \makeatother % There exist two functions $Q_1, Q_2: \mathbb{R}_+\to (0,\infty)$ of class \eqref{class:Q(r)} and two positive constants $C_{12}, r_0\in (0,\infty)$, such that % \begin{equation} \label{growth:q(x)} Q_1(|x|)\leq q(x)\leq Q_2(|x|)\leq C_{12}\, Q_1(|x|) \quad\mbox{for all } x\in \mathbb{R}^N \,, \end{equation} % together with % \begin{equation} \label{e:int-diff} \int_{r_0}^\infty ( Q_2(s) - Q_1(s) ) \int_{r_0}^s \exp \Big( - \int_r^s [ Q_1(t)^{1/2} + Q_2(t)^{1/2} ] \,\mathrm{d}t \Big) \,\mathrm{d}r \,\mathrm{d}s < \infty \,. \end{equation} % \end{enumerate} In fact, it suffices to assume inequalities \eqref{growth:q(x)} only for all $|x| = r > r_0$ with $r_0 > 0$ large enough, provided $q: \mathbb{R}^N\to (0,\infty)$ is a continuous function. Indeed, then one can find some extensions $\tilde{Q}_1, \tilde{Q}_2: \mathbb{R}_+\to (0,\infty)$ of class \eqref{class:Q(r)} of (the restrictions of) functions $Q_1, Q_2: [r_0 + 1, \infty)\to (0,\infty)$, respectively, from $[r_0 + 1, \infty)$ to $\mathbb{R}_+$ such that $\tilde{Q}_j(r) = Q_j(r)$ for $r\geq r_0 + 1$; $j=1,2$, and inequalities \eqref{growth:q(x)} hold for all $x\in \mathbb{R}^N$ with $\tilde{Q}_j$ in place of $Q_j$. \begin{remark}\label{rem-int-diff} \rm (a) Assuming \eqref{growth:q(x)}, we will show in the Appendix, {\S}\ref{ss:power-grow}, that the latter condition, \eqref{e:int-diff}, is satisfied in the following two cases: % \begin{itemize} % \item[{\rm (i)}] when $Q_2$ has at most power growth near infinity and the condition % \begin{equation} \int_{r_0}^\infty \left( Q_2(r)^{1/2} - Q_1(r)^{1/2} \right) \,{\rm d}r < \infty \quad\mbox{for some } 0 < r_0 < \infty \label{equiv:int-diff} \end{equation} % is valid; or % \item[{\rm (ii)}] when $Q_2$ has at most exponential power growth near infinity, i.e., $Q_2(r)\leq$\break $\gamma\cdot \exp( \beta r^{\alpha} )$ for all $r\geq r_0$, where $\alpha, \beta, \gamma > 0$ and $r_0 > 0$ are some constants, and % \begin{equation} \int_{r_0}^\infty \left( Q_2(r)^{1/2} - Q_1(r)^{1/2} \right) \left[ 1 + \log^+ ( Q_1(r)^{1/2} + Q_2(r)^{1/2} ) \right] \,\mathrm{d}r < \infty \,. \label{exp:int-diff} \end{equation} % \end{itemize} % Clearly, condition \eqref{exp:int-diff} is stronger than \eqref{equiv:int-diff}. {\rm (b)} With regard to Remark~{\rm \ref{rem-growth:q(r)}} and from the point of view of spectral theory, the case % \begin{equation} \label{growth:00$, we have also % \begin{equation} |f|\leq C\varphi \;\mbox{ in } \mathbb{R}^N \quad\Longrightarrow\quad |u|\leq C (\Lambda - \lambda)^{-1} \varphi \;\mbox{ in } \mathbb{R}^N , \label{e:max-princ_X} \end{equation} % by linearity. We denote by $\mathcal{K}\vert_{X}$ the restriction of $\mathcal{K} = (\mathcal{A} - \lambda I)^{-1}$ to the Banach space $X$ defined in \eqref{def:X}. Hence, $\mathcal{K}\vert_{X}$ is a bounded linear operator on $X$ with the operator norm $\leq (\Lambda - \lambda)^{-1}$, by \eqref{e:max-princ_X}. Clearly, $X$ is the dual space of the Lebesgue space $X^\odot = L^1(\mathbb{R}^N; \varphi \,{\rm d}x)$ with respect to the duality induced by the natural inner product on $L^2(\mathbb{R}^N)$. The embeddings \[ X\hookrightarrow L^2(\mathbb{R}^N)\hookrightarrow X^\odot \] are dense and continuous. Furthermore, $\mathcal{K}$ possesses a unique extension $\mathcal{K}\vert_{X^\odot}$ to a bounded linear operator on $X^\odot$ (by Lemma \ref{lem-Riesz-Thorin} below). Finally, it is obvious that $\mathcal{K}\vert_{X} : X\to X$ is the adjoint of $\mathcal{K}\vert_{X^\odot} : X^\odot\to X^\odot$. \subsection{Main theorems} \label{ss:Main-thm} Throughout this paragraph we assume that $q(x)$ is a potential that satisfies ~\eqref{hyp:growth:q(x)}. Under this hypothesis we are able to show the following \emph{ground\--state positivity } of the weak solution to the Schr\"odinger equation \eqref{e:Schroed_lam} in $X^\odot$. \begin{theorem}\label{thm-Positive-q(x)} Let \eqref{hyp:growth:q(x)} be satisfied and let $-\infty < \lambda < \Lambda$. Assume that $f\in X^\odot$ satisfies $f\geq 0$ almost everywhere and $f\not\equiv 0$ in~$\mathbb{R}^N$. Then the (unique) solution $u\in X^\odot$ to equation \eqref{e:Schroed_lam} (in the sense of distributions on $\mathbb{R}^N$) is given by $u = (\mathcal{A} - \lambda I)^{-1}\vert_{X^\odot} f$ and satisfies $u\geq c\varphi$ almost everywhere in $\mathbb{R}^N$, with some constant $c\equiv c(f) > 0$. % \end{theorem} In the literature, the inequality $u\geq c\varphi$ is often called briefly \emph{$\varphi$-positivity}. In {Protter} and {Weinberger} \cite[Chapt.~2, Theorem~10, p.~73]{ProtterWein}, a similar result is referred to as the \emph{generalized maximum principle}. This result has been established in {Alziary} and {Tak\'a\v{c}} \cite[Theorem 2.1, p.~284]{AlziaryTak} under slightly different growth hypotheses on the potentials $Q_1$ and $Q_2$ in conditions \eqref{growth:q(x)} and \eqref{e:int-diff}. In \cite[eq.~(2), p.~283]{AlziaryTak}, a closely related class~\eqref{class:Q(r)} is used where $Q(r)$ still satisfies a condition similar to \eqref{e:int-Q^-1/2}, namely, % \begin{math} \int_{r_0}^\infty Q(r)^{-\beta} \,{\rm d}r < \infty %\label{e:int-Q^-1/2} \end{math} % with some constant $\beta\in (0,1/2)$. Also the ``potential variation'' condition \eqref{e:int-diff} in our present work is somewhat different from that assumed in \cite[eq.~(5), p.~283]{AlziaryTak}. % Nevertheless, our proof of Theorem~\ref{thm-Positive-q(x)} follows similar steps as does the proof of Theorem 2.1 in \cite[pp.\ 289--290]{AlziaryTak}. Theorem~\ref{thm-Positive-q(x)} will be proved in Section~\ref{s:Proofs}, first for $q(x) = Q(|x|)$ of class~\eqref{class:Q(r)}, as Proposition~\ref{prop-Positive-Q(r)} in {\S}\ref{ss:Positive-Q(r)}, and then in its full generality in {\S}\ref{ss:Positive-q(x)}, after the proof of Theorem~\ref{thm-X-compact}, a part of which will be needed (stated below as Corollary~\ref{cor-X-compact}). The central result of this paper is the following compactness theorem. Indeed, it provides answers to some questions about the solution $u$ of problem \eqref{e:Schroed_lam}, such as $u\in X$ and its $\varphi$-positivity or $\varphi$-negativity. Moreover, it also guarantees that the spectrum of the operator $\mathcal{A} = - \Delta + q(x)\bullet$ is the same in each of the spaces $L^2(\mathbb{R}^N)$, $X$, and $X^\odot$. \begin{theorem}\label{thm-X-compact} Let \eqref{hyp:growth:q(x)} be satisfied. Then we have the following three statements for the resolvent $\mathcal{K} = (\mathcal{A} - \lambda I)^{-1}$ of $\mathcal{A}$ on $L^2(\mathbb{R}^N)$: % \begin{enumerate} % \item[{\rm (a)}] % If $-\infty < \lambda < \Lambda$ then both operators $\mathcal{K}\vert_{X}: X\to X$ and $\mathcal{K}\vert_{X^\odot}: X^\odot\to X^\odot$ are compact (and positive, see \eqref{e:max-princ_L^2}). % \item[{\rm (b)}] % If $\lambda\in \mathbb{C}$ is an eigenvalue of $\mathcal{A}$, that is, $\mathcal{A} v = \lambda v$ for some $v\in L^2(\mathbb{R}^N)$, $v\neq 0$, then $v\in X$ (${}\subset L^2(\mathbb{R}^N)\subset X^\odot$) and $\lambda\in \mathbb{R}$, $\lambda\geq \Lambda$. % \item[{\rm (c)}] % If $\lambda\in \mathbb{C}$ is \emph{not } an eigenvalue of $\mathcal{A}$, then the restriction $\mathcal{K}\vert_{X}$ of $\mathcal{K}$ to $X$ is a bounded linear operator from $X$ into itself and, moreover, $\mathcal{K}$ possesses a unique extension $\mathcal{K}\vert_{X^\odot}$ to a bounded linear operator from $X^\odot$ into itself. Again, both operators $\mathcal{K}\vert_{X}: X\to X$ and $\mathcal{K}\vert_{X^\odot}: X^\odot\to X^\odot$ are compact. % \end{enumerate} % \end{theorem} Part~(a) is the most difficult one to prove. Since $\mathcal{K}\vert_{X}: X\to X$ is compact {\it if and only if } $\mathcal{K}\vert_{X^\odot}: X^\odot\to X^\odot$ is compact, by Schauder's theorem ({Edwards} \cite[Corollary 9.2.3, p.~621]{Edwards} or {Yosida} \cite[Chapt.~X, Sect.~4, p.~282]{Yosida}), it suffices to prove that either of them is compact. Thus, our proof of Part~(a) begins with the compactness of the restriction of $\mathcal{K}\vert_{X}$ to (the corresponding subspace of) radially symmetric functions with $q(x) = Q(|x|)$ of class \eqref{class:Q(r)} and only for $\lambda < \Lambda$, see Lemma \ref{lem-X-compact}. So we may apply Schauder's theorem to get the compactness of the restriction of $\mathcal{K}\vert_{X^\odot}$ to radially symmetric functions with $q=Q$. Then we extend this result to $\mathcal{K}\vert_{X^\odot}$ on $X^\odot$ with $q=Q$ again, see Proposition \ref{prop-X-compact}. Finally, from there we derive that $\mathcal{K}\vert_{X^\odot}$ is compact for any $q(x)$ satifying \eqref{hyp:growth:q(x)}, first only for $\lambda < \Lambda$ and then for any $\lambda\in \mathbb{C}$ that is not an eigenvalue of $\mathcal{A}$, see Section~\ref{s:Proofs}, {\S}\ref{ss:Compact-q(x)}. Parts (b) and~(c) are proved immediately thereafter; they will be derived from Part~(a) by standard arguments based on the Riesz\--Schauder theory of compact linear operators ({Edwards} \cite[{\S}9.10, pp.\ 677--682]{Edwards} or {Yosida} \cite[Chapt.~X, Sect.~5, pp.\ 283--286]{Yosida}). As a direct consequence of Theorem \ref{thm-X-compact}, Part~(a), the following corollary establishes the equivalence of the ground states. \begin{corollary}\label{cor-X-compact} Let \eqref{hyp:growth:q(x)} be satisfied. Then the ground states $\varphi_{q}$, $\varphi_{Q_1}$, and $\varphi_{Q_2}$ corresponding to the potentials $q$, $Q_1$, and $Q_2$, respectively, are comparable, that is, there exist some constants $0 < \gamma_1\leq \gamma_2 < \infty$ such that % \begin{math} \gamma_1\varphi_{q} \leq \varphi_{Q_j} \leq \gamma_2\varphi_{q} \end{math} % in $\mathbb{R}^N$; $j=1,2$. Equivalently, we have $X_{q} = X_{Q_1} = X_{Q_2}$. % \end{corollary} A classical use of the compactness result, Theorem \ref{thm-X-compact}, Part~(c), is the {\it anti\--maximum principle } for the Schr\"odinger operator $\mathcal{A} = - \Delta + q(x)\bullet$ which complements the ground\--state positivity of Theorem~\ref{thm-Positive-q(x)}. \begin{theorem}\label{thm-Antimax-q(x)} Let \eqref{hyp:growth:q(x)} be satisfied and let $f\in X$ satisfy $\int_{\mathbb{R}^N} f\varphi \,{\rm d}x$ $> 0$. Then there exists a number $\delta\equiv \delta(f) > 0$ such that, for every $\lambda\in (\Lambda, \Lambda + \delta)$, the inequality $u\leq - c\varphi$ is valid a.e.\ in $\mathbb{R}^N$ with some constant $c\equiv c(f) > 0$. % \end{theorem} This important consequence of Theorem \ref{thm-X-compact} was presented also in {Alziary}, {Flec\-kinger}, and {Tak\'a\v{c}} \cite[Theorem 3.4]{AFT-3} for a class of potentials $q(x)$ satisfying different conditions (without radial symmetry). For a radially symmetric potential satisfying \eqref{hyp:growth:q(r)}, the anti-maximun principle has been obtained previously in {Alziary}, {Fleckinger}, and {Tak\'a\v{c}} \cite[Theorem 2.1, p.~128]{AFT-1} for $N=2$ and \cite[Theorem 2.1, p.~365]{AFT-2} for $N\geq 2$. Furthermore, in \cite{AFT-1, AFT-2} the function $f$ is assumed to be a ``sufficiently smooth'' perturbation of a radially symmetric function from~$X$. Theorem~\ref{thm-Antimax-q(x)} is an immediate consequence of the spectral decomposition of the resolvent of $\mathcal{A}$ as % \begin{equation} (\lambda I - \mathcal{A})^{-1} = (\lambda - \Lambda)^{-1} \mathcal{P} + \mathcal{H}(\lambda) \quad\mbox{for }\quad 0 < | \lambda - \Lambda | < \eta , \label{e:Laurent} \end{equation} % see, e.g., {Sweers} \cite[Theorem 3.2(ii), p.~259]{Sweers} or {Tak\'a\v{c}} \cite[Eq.~(6), p.~67]{Takac}. Here, $\lambda\in \mathbb{C}$, $\eta > 0$ is small enough, $\mathcal{P}$ denotes the spectral projection onto the eigenspace spanned by $\varphi$, and $\mathcal{H}(\lambda): L^2(\mathbb{R}^N)\to L^2(\mathbb{R}^N)$ is a holomorphic family of compact linear operators parametrized by $\lambda$ with $| \lambda - \Lambda | < \eta$. Moreover, $\mathcal{P}$ is selfadjoint and % \begin{math} \mathcal{P} \mathcal{H}(\lambda) = \mathcal{H}(\lambda) \mathcal{P} = 0 \end{math} % on $L^2(\mathbb{R}^N)$. Formula \eqref{e:Laurent} is used to prove the anti\--maximum principle also in \cite[Eq.\ (6), p.~124]{AFT-1} and \cite[Eq.\ (6), p.~361]{AFT-2}. % The main idea of the proof of Theorem~\ref{thm-Antimax-q(x)} is to show that each of the linear operators $\{ \mathcal{H}(\lambda): | \lambda - \lambda_1 | < \eta \}$ is bounded not only on $L^2(\mathbb{R}^N)$ but also on~$X$. Clearly, given the Neumann series expansion of $\mathcal{H}(\lambda)$ for $|\lambda - \Lambda| < \eta$, it suffices to show that the restriction $\mathcal{H}(\Lambda)\vert_X$ of $\mathcal{H}(\Lambda)$ to $X$ is a bounded linear operator on $X$. But this clearly follows from Theorem \ref{thm-X-compact}, Part~(c), with a help from formula~(6.32) in {Kato} \cite[Chapt.\ III, {\S}6.5, p.~180]{Kato} or formula~(1) in {Yosida} \cite[Chapt.\ VIII, Sect.~8, p.~228]{Yosida}. In various common versions of the anti\--maximum principle in a bounded domain $\Omega\subset \mathbb{R}^N$, $N\geq 1$, besides the assumption $0\leq f\not\equiv 0$ in $\Omega$, it is only assumed that $f\in L^p(\Omega)$ for some $p>N$ (cf.\ {Cl\'ement} and {Peletier} \cite[Theorem~1, p.~222]{ClemPel}, {Sweers} \cite{Sweers} or {Tak\'a\v{c}} \cite{Takac}). For $\Omega = \mathbb{R}^N$ the authors \cite[Example 4.1, pp.\ 377--379]{AFT-2} have constructed an example of a simple potential $q(r)$ and a function $f(r)$, both radially symmetric, $f\in L^2(\mathbb{R}^N)\setminus X$, and $0\leq f\not\equiv 0$ in $\mathbb{R}^N$, in which even the inequality $u\leq 0$ a.e.\ in $\mathbb{R}^N$ (weaker than the anti\--maximum principle of Theorem~\ref{thm-Antimax-q(x)}) is violated. More precisely, if $|\lambda - \Lambda| > 0$ is small enough, then even $u(r) > 0$ for every $r>0$ large enough. \section{Preliminary results} \label{s:Prelim} In {Alziary}, {Fleckinger}, and {Tak\'a\v{c}} \cite{AFT-3}, an important ingredient in comparing the ground states $\varphi_j(x)\equiv \varphi_{Q_j}(|x|)$ corresponding to the potentials $Q_j(|x|)$ ($j=1,2$) is an asymptotic formula due to {Hartman} and {Wintner} \cite[eq.\ (xxv), p.~49]{Hartman-Wint}. In contrast, here we take advantage of a generalized Titchmarsh' lemma (Lemma \ref{lem-Titchmarsh}) and a comparison result (Lemma \ref{lem-U1:U2}) to obtain the equivalence of the ground states, $\varphi_1$ and $\varphi_2$, under Hypothesis \eqref{hyp:growth:q(x)}. \subsection{Generalized Titchmarsh' lemma} \label{ss:Titchmarsh} Throughout this paragraph we assume that $U: [R,\infty)\to (0,\infty)$ is monotone increasing and continuous, where $0\leq R$ $< \infty$. Hence, $U'(r)\geq 0$ for a.e.\ $r\geq R$. We generalize Titchmarsh' lemma (cf.\ {Titchmarsh} \cite[Sect.~8.2, p.~165]{Titchmarsh}) used in {Alziary} and {Tak\'a\v{c}} \cite{AlziaryTak}, Lemma 3.2, p.~286. \begin{lemma}\label{lem-Titchmarsh} Assume that $f, f': [R,\infty)\to \mathbb{R}$ are locally absolutely continuous, $f(r) > 0$ for all $r\geq R$, and $f$ satisfies % \begin{equation} - f'' + U(r) f\leq 0 \quad\mbox{for a.e. } r\geq R . \label{ineq:f,f'} \end{equation} % If $f'\leq 0$ on $[R,\infty)$ then we must have % \begin{equation} - {f'}/{f} \geq U(r)^{1/2} \quad\mbox{for all } r\geq R . \label{ineq:g} \end{equation} % \end{lemma} \begin{proof} Upon the substitution $g = - (\log f)' = - f' / f$, where $g\geq 0$ on $[R,\infty)$, inequality \eqref{ineq:f,f'} is equivalent to % \begin{equation} g'\leq g^2 - U(r) \quad\mbox{for all } r\geq R . \label{ineq:g^2,g'} \end{equation} % This follows from % \begin{math} g'= {}- {f''}/{f} + ({f'}/{f})^2 = {}- {f''}/{f} + g^2 . \end{math} % Equivalently to \eqref{ineq:g}, we claim that $g\geq U^{1/2}$ on $[R,\infty)$. Indeed, if on the contrary \[ \delta\stackrel{{\rm def}}{=} g(r_0)^2 - U(r_0) < 0 \quad\mbox{for some } r_0\geq R , \] then from \eqref{ineq:g^2,g'} and the continuity of $g$ and $U$ at $r_0$ we deduce that there exists some $r_1 > r_0$ such that also \[ g'(r)\leq g(r)^2 - U(r) < 0 \quad\mbox{for a.e. } r\in [r_0, r_1) . \] Consequently, with regard to $U(r)\geq U(r_0)$ for $r\geq r_0$, we get % \begin{equation*} g'(r)\leq g(r)^2 - U(r) < g(r_0)^2 - U(r)\leq g(r_0)^2 - U(r_0) = \delta < 0 \end{equation*} % for a.e.\ $r\in (r_0, r_1]$. Hence, we may take $r_1 > r_0$ arbitrarily large as long as $g(r_1)\geq 0$ holds; let us choose $r_1 > r_0$ so large that $g(r_1) = 0$. Then \[ g'(r_1)\leq g(r_1)^2 - U(r_1) = - U(r_1) < 0 \] which implies $g(r_1 + s) < 0$ for every $s>0$ small enough and thus contradicts our hypothesis $g\geq 0$ on $[R,\infty)$. Hence, we have proved $g(r)^2 - U(r)\geq 0$ for all $r\geq R$, which entails $-f'/f = g\geq U^{1/2}$ on $[R,\infty)$. \end{proof} \begin{corollary}\label{cor-Titchmarsh} In the situation of {\rm Lemma~\ref{lem-Titchmarsh}} above, the function $f$ is decreasing, convex, and satisfies $f(s)\searrow 0$ and $f'(s)\nearrow 0$ as $s\nearrow \infty$. Moreover, we have % \begin{equation} \frac{f(s)}{f(r)} \leq \exp\Big( - \int_r^s U(t)^{1/2} \,{\rm d}t \Big) \quad\mbox{whenever }\, R\leq r\leq s < \infty . % \label{ineq:f(s)/f(r)} \end{equation} % \end{corollary} \begin{proof} First, we integrate inequality \eqref{ineq:g} over the interval $[r,s]$ to get \eqref{ineq:f(s)/f(r)}. Since $f(s) > 0$ for all $s\geq R$, and $U$ is monotone increasing, inequality \eqref{ineq:f(s)/f(r)} forces $f(s)\searrow 0$ as $s\nearrow \infty$. Next, \eqref{ineq:f,f'} guarantees that $f$ is convex. It follows that $f'(s)\nearrow 0$ as $s\nearrow \infty$. \end{proof} \subsection{Asymptotic equivalence of solutions} \label{ss:phi_1/phi_2} In this paragraph we compare positive solutions of homogeneous Schr\"odinger equations (or inequalities) in $[R,\infty)$ with different potentials (for $N=1$). % We start with a comparison result proved in {M.~Hoffmann\--Ostenhof} et al.\ \cite[Lemma 3.2, p.~348]{HoffSwetina}. \begin{lemma}\label{lem-U1:U2} Let $U_1, U_2: [R,\infty)\to (0,\infty)$ be two continuous potentials satisfying $0 < \mbox{\rm const} \leq U_1\leq U_2$ for $r\geq R$, where $0 < R < \infty$. Let all $f_1, f_2, f_1', f_2'\in L^2(R,\infty)$ be locally absolutely continuous in $[R,\infty)$, and let also $f_1 > 0$ and $f_2 > 0$ for $r\geq R$. Finally, assume that \[ -f_1'' + U_1(r) f_1\geq 0 \quad\mbox{and }\quad -f_2'' + U_2(r) f_2\leq 0 \quad\mbox{for almost every }\ r > R . \] Then we have \[ \frac{f_1 }{f_2}\geq \frac{f_1(R)}{f_2(R)} \;\mbox{ and }\; \frac{f_1'}{f_1}\geq \frac{f_2' }{f_2 } \quad\mbox{for every }\ r\geq R . \] % \end{lemma} In the next proposition we give a sufficient upper bound for the perturbation $Q_2(r) - Q_1(r)$ in \eqref{hyp:growth:q(x)} that guarantees that all ground states $\varphi_{q}$, $\varphi_{Q_1}$, and $\varphi_{Q_2}$ are comparable (see {\S}\ref{ss:Compact-q(x)}). \begin{proposition}\label{prop-Titchmarsh} Let $U_1, U_2: [R,\infty)\to (0,\infty)$ be monotone increasing and continuous, where $0\leq R < \infty$. Assume that $f_j, f_j': [R,\infty)\to \mathbb{R}$ are locally absolutely continuous; $j=1,2$, $f_j(r) > 0$ and $f_j'(r)\leq 0$ for all $r\geq R$, and $f_j$ satisfies % \begin{equation} - f_j'' + U_j(r) f_j = 0 \quad\mbox{for a.e. } r\geq R . \label{eq:f_j,f_j'} \end{equation} % Then for all $r\geq R$ we have % \begin{equation} \label{ineq:log(f_1/f_2)} \begin{split} & \left\vert \frac{{\rm d}}{\mathrm{d}r} \left( \log\, \frac{f_1(r)}{f_2(r)} \right) \right\vert \\ &\leq \int_r^\infty | U_1(s) - U_2(s) |\, \exp \Big( - \int_r^s [ U_1(t)^{1/2} + U_2(t)^{1/2} ] \,\mathrm{d}t \Big) \,\mathrm{d}s \,. \end{split} \end{equation} % \end{proposition} \begin{proof} Employing \eqref{eq:f_j,f_j'} we compute \[ \frac{{\rm d}}{{\rm d}r} \left[ f_1 f_2\left( \frac{f_1'}{f_1} - \frac{f_2'}{f_2} \right) \right] = f_1'' f_2 - f_1 f_2'' = ( U_1(r) - U_2(r) )\, f_1 f_2 . \] Since $f_1'(r) f_2(r)\to 0$ and $f_1(r) f_2'(r)\to 0$ as $r\to \infty$, by Corollary~\ref{cor-Titchmarsh}, integration yields \[ - f_1(r) f_2(r)\left( \frac{f_1'}{f_1} - \frac{f_2'}{f_2} \right) = \int_r^\infty ( U_1(s) - U_2(s) )\, f_1(s) f_2(s) \,{\rm d}s , \] i.e., % \begin{equation*} \frac{\mathrm{d}}{\mathrm{d}r} \left( \log\, \frac{f_1}{f_2} \right) = {}- \frac{1}{ f_1(r) f_2(r) } \int_r^\infty ( U_1(s) - U_2(s) )\, f_1(s) f_2(s) \,\mathrm{d}s \end{equation*} % for all $r\geq R$. Consequently, % \begin{equation*} \left\vert \frac{\mathrm{d}}{\mathrm{d}r} \left( \log\, \frac{f_1(r)}{f_2(r)} \right) \right\vert \leq \frac{1}{ f_1(r) f_2(r) } \int_r^\infty | U_1(s) - U_2(s) |\, f_1(s) f_2(s) \,\mathrm{d}s \end{equation*} % for all $r\geq R$. Finally, we apply inequality \eqref{ineq:f(s)/f(r)} with $U = U_j$ and $f = f_j$; $j=1,2$, to the right\--hand side of the inequality above to get \eqref{ineq:log(f_1/f_2)}. \end{proof} \begin{remark}\label{rem-Titchmarsh}\rm In the situation of {\rm Proposition~\ref{prop-Titchmarsh}} above, the functions $f_1(r)$ and $f_2(r)$ are \emph{asymptotically equivalent near infinity }, meaning that the ratios ${f_1(r)}/{f_2(r)}$ and ${f_2(r)}/{f_1(r)}$ stay bounded as $r\to \infty$, provided the following condition is satisfied: % \begin{equation} \label{e:int-diff:U_2-U_1} \begin{split} & \int_R^\infty \int_r^\infty | U_1(s) - U_2(s) |\, \exp\Big( - \int_r^s [ U_1(t)^{1/2} + U_2(t)^{1/2} ] \,{\rm d}t \Big) \,{\rm d}s \,{\rm d}r = \\ & \int_R^\infty | U_1(s) - U_2(s) | \int_R^s \exp\Big( - \int_r^s [ U_1(t)^{1/2} + U_2(t)^{1/2} ] \,{\rm d}t \Big) \,{\rm d}r \,{\rm d}s < \infty \,. \end{split} \end{equation} % This claim follows by integrating \eqref{ineq:log(f_1/f_2)} with respect to $r\in [R,\infty)$. Note that condition \eqref{e:int-diff:U_2-U_1} corresponds to \eqref{e:int-diff} for $Q_1$ and $Q_2$. % \end{remark} Simple sufficient conditions for $U_1$ and $U_2$ are given in the Appendix, {\S}\ref{ss:power-grow}, which guarantee that \eqref{e:int-diff:U_2-U_1} is satisfied. \section{Two known compactness results} \label{s:AFT-compact} We state in this section two compactness results obtained in {Alziary}, {Fleckin\-ger}, and {Tak\'a\v{c}} \cite[Section~6]{AFT-3} which are essential for the proof of our main Theorem \ref{thm-X-compact}. We need local compactness from Proposition \ref{prop-X^o-B_R} together with a compactness result by comparison of two potentials from Proposition \ref{prop-q_1 0$ throughout $\mathbb{R}^N$ and $\|\varphi_j\|_{L^2(\mathbb{R}^N)} = 1$. Finally, we write $X_j = X_{q_j}$ and $X_j^\odot = L^1(\mathbb{R}^N; \varphi_j \,{\rm d}x)$. The following comparison result is natural (and holds without any growth conditions other than \eqref{cond:q_0}). \begin{proposition}\label{prop-q_1 r_0$, is monotone increasing, $\int_{r_0}^\infty P(r)^{-1} \,{\rm d}r$ $< \infty$, and % \begin{equation} p(r)\stackrel{{\rm def}}{=} {}- \frac{N-1}{r} - 2 (\log\varphi(r))' \geq P(r) > 0 \quad\mbox{holds for all } r > r_0 . \label{growth:phi(r)} \end{equation} % \end{lemma} \begin{proof} Using the radial Schr\"odinger equation for $\varphi$, % \begin{equation*} - \varphi''(r) - \frac{N-1}{r}\, \varphi'(r) + q(r) \varphi(r) = \Lambda \varphi(r) \quad\mbox{for } r\geq R , \end{equation*} % we observe that the function $v(r) = r^{(N-1)/2} \varphi(r) > 0$ must satisfy % \begin{equation*} - v''(r) + \left( q(r) - \Lambda + \frac{(N-1)(N-3)}{4 r^2} \right) v(r) = 0 \quad\mbox{for } r\geq R . \end{equation*} % It follows from \eqref{growth:q(r)} that % \begin{equation*} - v''(r) + Q(r) v(r)\leq 0 \quad\mbox{for } r\geq R . \end{equation*} % Next, we claim that $v'(r)\leq 0$ for all $r\geq R$. Indeed, since $v''(r)\geq Q(r) v(r)\geq 0$, the derivative $v'$ is nondecreasing on $[R,\infty)$. Therefore, if $v'(r_0) > 0$ for some $r_0\geq R$, then $v'(r)\geq v'(r_0) > 0$ for every $r\geq r_0$, which contradicts $\int_{r_0}^\infty v(r)\, r^{(N-1)/2} \,{\rm d}r < \infty$. Finally, we may apply the generalized Titchmarsh' lemma to conclude that the function $g = - (\log v)' = - v' / v$ satisfies $g(r)\geq Q(r)^{1/2}$ for all $r\geq R$. We compute \[ g(r) = {}- \frac{{\rm d}}{{\rm d}r} \log \left( r^{(N-1)/2} \varphi(r) \right) = {}- \frac{N-1}{2r} - \frac{{\rm d}}{{\rm d}r}\, \log \varphi(r) = \textstyle\frac12 \, p(r) \] with $g(r)\geq Q(r)^{1/2}$ for all $r\geq R$. This proves \eqref{growth:phi(r)}. The remaining claims follow from the properties of class \eqref{class:Q(r)}. \end{proof} We prove Lemma \ref{lem-X-compact} directly using Arzel\`a\--Ascoli's compactness criterion for continuous functions on the one point compactification $\mathbb{R}_+^* = \mathbb{R}_+\cup \{\infty\}$ of $\mathbb{R}_+$. The metric on $\mathbb{R}_+^*$ is defined by % \begin{equation*} d(x,y)\stackrel{{\rm def}}{=} \begin{cases} \frac{|x-y|}{1 + |x-y|} & \mbox{for } x,y\in \mathbb{R}_+ ; \\ 1 & \mbox{for } 0\leq x < y = \infty \;\mbox{ or }\; 0\leq y < x = \infty ;\\ 0 & \mbox {for }\; x = y = \infty . \end{cases} \end{equation*} % We denote by $C(\mathbb{R}_+^*)$ the Banach space of all continuous functions on the compact metric space $\mathbb{R}_+^*$ endowed with the supremum norm from $L^\infty(\mathbb{R}_+)$. \begin{proof}[Proof of Lemma~\ref{lem-X-compact}] Given $f,u\in X_\mathrm{rad}$, $u = \mathcal{K} f$ is equivalent with the ordinary differential equation % \begin{equation*} - u''(r) - \frac{N-1}{r}\, u'(r) + q(r) u(r) = \lambda u(r) + f(r) \quad\mbox{for } 0 < r < \infty %\label{eq:u=Kf} \end{equation*} % supplemented by the conditions % \begin{equation*} \lim_{r\to 0+} u'(r) = 0 \quad\mbox{and }\quad \sup_{0 < r < \infty} \genfrac{|}{|}{}0{u(r)}{\varphi(r)} \leq (\Lambda - \lambda)^{-1} \cdot \sup_{0 < r < \infty} \genfrac{|}{|}{}0{f(r)}{\varphi(r)} \,. %\label{bc:u=Kf} \end{equation*} % Clearly, the former one is a boundary condition at zero that follows from the radial symmetry, whereas the latter one follows from the weak maximum principle. Substituting $g = f / \varphi$ and $v = u / \varphi$, combined with % \begin{equation*} - \varphi''(r) - \frac{N-1}{r}\, \varphi'(r) + q(r) \varphi(r) = \Lambda \varphi(r) \quad\mbox{for } 0 < r < \infty , \end{equation*} % we have equivalently % \begin{equation} - v''(r) - \frac{N-1}{r}\, v'(r) - 2 (\log\varphi(r))'\, v'(r) + (\Lambda - \lambda) v(r) = g(r) \quad\mbox{for } 0 < r < \infty \label{eq:v} \end{equation} % subject to the conditions % \begin{equation} \lim_{r\to 0+} v'(r) = 0 \quad\mbox{and }\quad \sup_{0 < r < \infty} |v(r)| \leq (\Lambda - \lambda)^{-1} \cdot \sup_{0 < r < \infty} |g(r)| \,. \label{cond:v} \end{equation} % Then $\mathcal{K}\vert_{X_\mathrm{rad}}$ is compact on $X_\mathrm{rad}$ if and only if the linear operator $\mathcal{K}_{\varphi} : L^\infty(\mathbb{R}_+)\to L^\infty(\mathbb{R}_+)$, defined by \[ \mathcal{K}_{\varphi} g\stackrel{{\rm def}}{=} v = \varphi^{-1}\cdot \mathcal{K} (g\varphi) \quad\mbox{for } g\in L^\infty(\mathbb{R}_+) , \] is compact. We will apply Arzel\`a\--Ascoli's compactness criterion in the Banach space $C(\mathbb{R}_+^*)$ in order to show that the image % \begin{math} \mathcal{K}_{\varphi} \left( \overline{\mathcal{B}}_{ L^\infty(\mathbb{R}_+) } \right) \end{math} % of the unit ball \[ \overline{\mathcal{B}}_{ L^\infty(\mathbb{R}_+) } = \left\{ g\in L^\infty(\mathbb{R}_+): \| g\|_{ L^\infty(\mathbb{R}_+) } \leq 1 \right\} \] has compact closure in $C(\mathbb{R}_+^*)$. Since $L^\infty(\mathbb{R}_+)$ is a Banach lattice, it suffices to show that % \begin{math} \mathcal{K}_{\varphi} \Big( \overline{\mathcal{B}}_{ L^\infty(\mathbb{R}_+) }^+ \Big) \end{math} % has compact closure in $C(\mathbb{R}_+^*)$, where \[ \overline{\mathcal{B}}_{ L^\infty(\mathbb{R}_+) }^+ = \left\{ g\in \overline{\mathcal{B}}_{ L^\infty(\mathbb{R}_+) } : g\geq 0 \,\mbox{ in }\, \mathbb{R}_+ \right\} . \] Clearly, the function $v$ from \eqref{eq:v} and \eqref{cond:v} above satisfies $v\in C^1(\mathbb{R}_+)$; we will show also $v\in C(\mathbb{R}_+^*)$. Therefore, we need to show that the linear operator % \begin{equation*} \mathcal{K}_{\varphi} : L^\infty(\mathbb{R}_+)\to C(\mathbb{R}_+^*)\subset L^\infty(\mathbb{R}_+) \end{equation*} % is compact. So let $g\in L^\infty(\mathbb{R}_+)$ be arbitrary with $0\leq g(r)\leq 1$ for $r\in \mathbb{R}_+$. Hence, $v = \mathcal{K}_{\varphi} g$ satisfies $v\in C^1(\mathbb{R}_+)$ and also $0\leq v(r)\leq (\Lambda - \lambda)^{-1}$, by \eqref{e:max-princ_X}. It follows that the function % \begin{equation*} g^\sharp\stackrel{{\rm def}}{=} g - (\Lambda - \lambda) v \end{equation*} % satisfies $-1\leq g^\sharp\leq 1$, and the derivative $w\stackrel{{\rm def}}{=} v'$ verifies the ordinary differential equation % \begin{equation} - w'(r) - \frac{N-1}{r}\, w(r) - 2 (\log\varphi(r))'\, w(r) = g^\sharp(r) \quad\mbox{for } 0 < r < \infty \label{eq:w} \end{equation} % subject to the conditions % \[ \lim_{r\to 0+} w(r) = 0 \quad\mbox{and }\quad \sup_{0 < r < \infty} \left| \int_0^r w(s) \,{\rm d}s \right| \leq (\Lambda - \lambda)^{-1} \,. %\label{cond:w} \] % The latter condition has been obtained from % \[ \int_0^r w(s) \,{\rm d}s = v(r) - v(0) \quad\mbox{with }\quad 0\leq v(r)\leq (\Lambda - \lambda)^{-1} \] % for all $r\geq 0$. Since $w$ is continuous, this condition implies that there exists a sequence $\{ r_n\}_{n=1}^\infty \subset \mathbb{R}_+$ such that $r_n\to \infty$ and $w(r_n)\to 0$ as $n\to \infty$. The differential equation \eqref{eq:w} is equivalent to % \begin{equation*} - \frac{{\rm d}}{{\rm d}r} \left( r^{N-1} \varphi(r)^2\, w(r) \right) = r^{N-1} \varphi(r)^2\, g^\sharp(r) \quad\mbox{for } 0 < r < \infty . %\label{eq:w:equiv} \end{equation*} % After integration, we thus arrive at % \[ r^{N-1} \varphi(r)^2\, w(r) - s^{N-1} \varphi(s)^2\, w(s) = \int_{r}^{s} t^{N-1} \varphi(t)^2\, g^\sharp(t) \,{\rm d}t %\label{eq:w:int_r^s} \] % whenever $0\leq r,s < \infty$. Applying $\lim_{s\to 0+} w(s) = 0$ we obtain % \begin{equation} r^{N-1} \varphi(r)^2\, w(r) = - \int_{0}^{r} t^{N-1} \varphi(t)^2\, g^\sharp(t) \,{\rm d}t \quad\mbox{for all } r\geq 0 . \label{eq:w:int_0^r} \end{equation} % Taking $s = r_n$ and letting $n\to \infty$ we obtain also % \begin{equation} r^{N-1} \varphi(r)^2\, w(r) = \int_{r}^{\infty} t^{N-1} \varphi(t)^2\, g^\sharp(t) \,{\rm d}t \quad\mbox{for all } r\geq 0 . \label{eq:w:int_r} \end{equation} % Here we have used the facts that $s^{N-1} \varphi(s)^2 \to 0$ as $s\to \infty$ together with $r_n\to \infty$ and $w(r_n)\to 0$ as $n\to \infty$. Recall the normalization % \begin{math} \int_0^\infty \varphi(t)^2\, t^{N-1} \,{\rm d}t = \sigma_{N-1}^{-1} , \end{math} % where $\sigma_{N-1}$ stands for the surface area of the unit sphere in $\mathbb{R}^N$. Below we will take advantage of formulas \eqref{eq:w:int_0^r} and \eqref{eq:w:int_r} to estimate $|w(r)|$ as $r\to 0+$ and $r\to \infty$, respectively. Because of $|g^\sharp|\leq 1$, eq.~\eqref{eq:w:int_0^r} yields $|w|\leq w^\sharp_0$ where $w^\sharp_0: \mathbb{R}_+\to \mathbb{R}_+$ is the function defined by $w^\sharp_0(0) = 0$ and % \[ r^{N-1} \varphi(r)^2\, w^\sharp_0(r) = \int_{0}^{r} t^{N-1} \varphi(t)^2 \,{\rm d}t \quad\mbox{for all } r > 0 . %\label{eq:^w_n:int_0^r} \] % Using $\lim_{r\to 0+} \varphi(r) = \varphi(0) > 0$ we conclude that % \begin{equation*} %\label{est:^w_0:int_0^r} \begin{split} \lim_{r\to 0+} \,\frac{w^\sharp_0(r)}{r} & = \lim_{r\to 0+} \frac{1}{r} \int_{0}^{r} \genfrac{(}{)}{}0{t}{r}^{N-1} \genfrac{(}{)}{}0{\varphi(t)}{\varphi(r)}^2 \,{\rm d}t \\ & = \lim_{r\to 0+} \frac{1}{r} \int_{0}^{r} \genfrac{(}{)}{}0{t}{r}^{N-1} \,{\rm d}t = \frac{1}{N}\, . \end{split} \end{equation*} % Because of $|g^\sharp|\leq 1$, eq.~\eqref{eq:w:int_r} yields $|w|\leq w^\sharp_\infty$ where $w^\sharp_\infty: (0,\infty)\to \mathbb{R}_+$ is the function defined by % \begin{equation} w^\sharp_\infty(r) = r^{-(N-1)} \varphi(r)^{-2} \int_{r}^{\infty} t^{N-1} \varphi(t)^2 \,{\rm d}t \quad\mbox{for all } r > 0 . \label{eq:^w_infty:int_r} \end{equation} % Next, we wish to show that $w^\sharp_\infty(r)\leq P(r)^{-1}$ holds for all $r > r_0$, where $P(r) = 2\, Q(r)^{1/2}$. To this end, notice first that eq.~\eqref{eq:^w_infty:int_r} is equivalent with % \begin{equation*} w^\sharp_{\infty}(r) = \int_{r}^{\infty} \exp\Big( - \int_r^s p(t) \,{\rm d}t \Big) \,{\rm d}s , \quad r > r_0 , \end{equation*} % where $p(t)$ is given by formula \eqref{growth:phi(r)}. In particular, the function $w^\sharp_\infty(r)$ satisfies the differential equation % \begin{equation*} - \frac{{\rm d}}{{\rm d}r}\, w^\sharp_{\infty}(r) + p(r)\, w^\sharp_\infty(r) = 1 \quad\mbox{for } r_0 < r < \infty . \end{equation*} % By Lemma \ref{lem-growth:phi(r)}, we have $p(t)\geq P(t)\geq P(r)$ whenever $r_0 < r\leq t < \infty$. Hence, we can estimate % \begin{equation*} \begin{split} w^\sharp_{\infty}(r) & \leq \int_r^{\infty} \exp\Big( - \int_r^s P(r) \,{\rm d}t \Big) \,{\rm d}s \\ & = \int_r^{\infty} e^{ - P(r)\, (s-r) } \,{\rm d}s = \int_0^{\infty} e^{ - P(r)\, s } \,{\rm d}s = P(r)^{-1} . \end{split} \end{equation*} % To summarize our estimates for the functions $w^\sharp_0: \mathbb{R}_+\to \mathbb{R}_+$ and $w^\sharp_\infty: (0,\infty)\to \mathbb{R}_+$ in the inequalities $|w|\leq w^\sharp_0$ for $r\geq 0$ and $|w|\leq w^\sharp_\infty$ for $r > r_0$, we observe that both functions $w^\sharp_0$ and $w^\sharp_\infty$ are continuously differentiable and satisfy the estimates % \begin{gather} |w(r)|\leq w^\sharp_0(r)\leq C\, r \quad\mbox{for }\; 0\leq r\leq r_0 , \label{est:^w_0:rr_0} \end{gather} % where $C>0$ is a constant. Recall that $\int_{r_0}^\infty P(r)^{-1} \,{\rm d}r < \infty$, by condition \eqref{e:int-Q^-1/2}. Consequently, for $g$ ranging over $L^\infty(\mathbb{R}_+)$ with $0\leq g\leq 1$ in $\mathbb{R}_+$, the set of functions $v = \mathcal{K}_{\varphi} g\in C^1(\mathbb{R}_+)$ defined above is {\it uniformly equicontinuous } on the compact metric space $\mathbb{R}_+^*$, thanks to % \[ v(r) = v(0) + \int_0^r w(s) \,{\rm d}s = v(\infty) - \int_r^\infty w(s) \,{\rm d}s \quad\mbox{for } 0\leq r < \infty . \] % The limit $v(\infty) = \lim_{r\to \infty} v(r) \in \mathbb{R}$ exists by \eqref{est:^w_infty:r>r_0}. Furthermore, owing to $0\leq v(r)\leq (\Lambda - \lambda)^{-1}$ for $r\in \mathbb{R}_+$, this set is also {\it uniformly bounded } on $\mathbb{R}_+^*$. Thus, by Arzel\`a\--Ascoli's compactness criterion, the set % \begin{math} \mathcal{K}_{\varphi} \Big( \overline{\mathcal{B}}_{ L^\infty(\mathbb{R}_+) }^+ \Big) \end{math} % has compact closure in $C(\mathbb{R}_+^*)$. We have proved that the linear operator $\mathcal{K}\vert_{X_\mathrm{rad}}$ is compact on $X_\mathrm{rad}$ and, moreover, its image satisfies $\mathcal{K}(X_\mathrm{rad}) \subset C(\mathbb{R}_+^*)$. \end{proof} \subsection{Compactness on the entire space $X$} \label{ss:Compact-gen} We keep the assumption $q(x) = Q(|x|)$ for all $x\in \mathbb{R}^N$. Recall $\lambda < \Lambda$ and $\mathcal{K} = (\mathcal{A} - \lambda I)^{-1}$ on $L^2(\mathbb{R}^N)$. This time we will show first that the operator $\mathcal{K}\vert_{X^\odot}$ is compact on $X^\odot$. We derive this result from the compactness of its restriction $\mathcal{K}\vert_{ X_\mathrm{rad}^\odot }$ to $X_\mathrm{rad}^\odot$ which we have already established in the previous paragraph. To prove the compactness of $\mathcal{K}\vert_{X^\odot}$, we will apply the well\--known compactness criterion of Fr\'echet and Kolmogorov in the Lebesgue space $X^\odot =$\break $L^1(\mathbb{R}^N; \varphi \,{\rm d}x)$; see {Edwards} \cite[Theorem 4.20.1, p.~269]{Edwards} or {Yosida} \cite[Chapt.~X, Sect.~1, p.~275]{Yosida}. We denote by \[ \overline{\mathcal{B}}_{X^\odot} = \left\{ f\in X^\odot: \| f\|_{X^\odot} \leq 1 \right\} \] the closed unit ball centered at the origin in the Banach lattice $X^\odot = L^1(\mathbb{R}^N; \varphi \,\mathrm{d}x)$. \begin{lemma}\label{lem-X^o-infty} Given any $\varepsilon > 0$, there exists a number $R\equiv R(\varepsilon)\in (0,\infty)$ such that for every $f\in \overline{\mathcal{B}}_{X^\odot}$ and $u = \mathcal{K}\vert_{X^\odot} f$ we have % \begin{equation} \int_{|x|\geq R} |u(x)|\, \varphi(|x|) \,{\rm d}x \leq \varepsilon . \label{est:int_|u| 0$. % \end{proposition} This proposition is proved in {Alziary} and {Tak\'a\v{c}} \cite[Theorem 2.1, p.~284]{AlziaryTak} under a slightly different growth hypothesis on the monotone increasing potential $q(x)\equiv q(r)$. \begin{proof} The proof of our proposition follows the same pattern as does the proof of Theorem 2.1 in \cite[pp.\ 289--290]{AlziaryTak}. We leave the details to the reader. \end{proof} The results of the previous section ({\S}\ref{ss:compact-q_1 r_0 ;\ j=1,2 , \label{def:V_j(r)} \end{equation} % where $0 < r_0 < \infty$ is large enough, such that $| {(N-1)(N-3)} | / {4 r_0^2} \leq 1$ and $V_j(r) > 0$ for all $r > r_0$. Consequently, \[ Q_j(r) - \Lambda - 1\leq V_j(r)\leq Q_j(r) - \Lambda + 1 \quad\mbox{for all } r > r_0 ;\ j=1,2 . \] Therefore, we may apply first Remarks \ref{rem-int-diff}, Part~{\rm (b)}, and \ref{rem-Titchmarsh}, and then Lemma \ref{lem-U1:U2} to conclude that $\varphi_{Q_j}(r)$ is comparable with any positive solution $f_j(r)$ of % \begin{equation*} - f_j'' + Q_j(r) f_j = 0 \quad\mbox{for a.e. } r > r_0 , %\label{eq:f_j,f_j'} \end{equation*} % such that $f_j(r)\to 0$ as $r\to \infty$. The functions $f_1$ and $f_2$ are comparable by results from paragraph {\S}\ref{ss:phi_1/phi_2}, Proposition \ref{lem-Titchmarsh} and Remark \ref{rem-Titchmarsh} for $U_j = Q_j$; $j=1,2$. Thus, we have obtained % \begin{equation} \label{growth:phi_1/phi_2} 0 < c'\leq \frac{\varphi_{Q_1}(r)}{\varphi_{Q_2}(r)} \leq c''< \infty \quad\mbox{for all } r > r_0 , \end{equation} % where $c'$ and $c''$ are some constants. This formula yields $\sup_{\mathbb{R}^N} (\varphi_{Q_1} / \varphi_{Q_2}) < \infty$ or, equivalently, $X_{Q_1}\hookrightarrow X_{Q_2}$ is a continuous embedding. Finally, let us rewrite the equation $\mathcal{A}_{q} \varphi_{q} = \Lambda_{q} \varphi_{q}$ for $\varphi_{q}\in X_{Q_1} = X_{Q_2}$ as % \begin{equation*} - \Delta\varphi_{q} + Q_2(|x|)\varphi_{q} = f(x) \quad\mbox{in }\, X_{Q_2}^\odot , %\label{Q_2(r):Schroed_0} \end{equation*} % where % \begin{equation*} f(x) = \left[ Q_2(|x|) - q(x) + \Lambda_{q} \right] \varphi_{q}(x) \geq \Lambda_{q} \varphi_{q}(x) > 0 ,\quad x\in \mathbb{R}^N , \end{equation*} % by condition \eqref{growth:q(x)}. Notice that % \begin{equation*} \sup_{\mathbb{R}^N} (\varphi_{q} / \varphi_{Q_1}) < \infty , \quad \sup_{\mathbb{R}^N} (\varphi_{Q_2} / \varphi_{Q_1}) < \infty , \quad\mbox{and}\quad \sup_{\mathbb{R}^N} (\varphi_{Q_1} / \varphi_{Q_2}) < \infty , \end{equation*} % combined with $\int_{\mathbb{R}^N} Q_2\, \varphi_{Q_2}^2 \,{\rm d}x < \infty$, yield $\int_{\mathbb{R}^N} Q_2\, \varphi_{q}\, \varphi_{Q_2} \,{\rm d}x < \infty$, that is, % \begin{math} Q_2\, \varphi_{q}\in X_{Q_2}^\odot = L^1(\mathbb{R}^N; \varphi_{Q_2} \,{\rm d}x) . \end{math} % Consequently, also $(Q_2 - q)\varphi_{q}\in X_{Q_2}^\odot$ which guarantees $f\in X_{Q_2}^\odot$. We apply Proposition~\ref{prop-Positive-Q(r)} with $Q=Q_2$ and $\lambda = 0 < \Lambda_Q$ to conclude that $\inf_{\mathbb{R}^N} (\varphi_{q} / \varphi_{Q_2}) > 0$ or, equivalently, $X_{Q_2}\hookrightarrow X_{q}$ is a continuous embedding. Summarizing the results proved in this section for $\varphi_{q}$, $\varphi_{Q_1}$, and $\varphi_{Q_2}$, we arrive at $X_{q} = X_{Q_1} = X_{Q_2}$, i.e., % \begin{math} \gamma_1\varphi_{q} \leq \varphi_{Q_1}, \varphi_{Q_2} \leq \gamma_2\varphi_{q} \end{math} % everywhere in $\mathbb{R}^N$, where $0 < \gamma_1\leq \gamma_2 < \infty$ are some constants. As we already know that the restriction $(\mathcal{A}_{q} - \lambda I)^{-1}\vert_{X_{Q_1}}$ to $X_{Q_1}$ is compact, Part~(a) follows immediately. {\rm Part~(b)}: $\;$ In the remaining part of the proof we abbreviate $\mathcal{A} = \mathcal{A}_{q}$, $\Lambda = \Lambda_{q}$, and $X = X_{q}$. Let $\lambda\in \mathbb{C}$ be an eigenvalue of $\mathcal{A}$, that is, $\mathcal{A} v = \lambda v$ for some $v\in L^2(\mathbb{R}^N)$, $v\neq 0$. Since $\mathcal{A}$ is positive definite and selfadjoint on $L^2(\Omega)$, its inverse $\mathcal{A}^{-1}$ is bounded on $L^2(\mathbb{R}^N)$. Property \eqref{cond:q_0} implies that $\mathcal{A}^{-1}$ is also compact. Consequently, $\lambda\in \mathbb{R}$ and $\lambda\geq \Lambda > 0$. Given $v\in L^2(\mathbb{R}^N)$, $v\neq 0$, it follows that equation $\mathcal{A} v = \lambda v$ is equivalent with $\mathcal{A}^{-1} v = \lambda^{-1} v$. By Part~(a), also the restriction $\mathcal{A}^{-1}\vert_{X}$ to $X$ is compact. Now we can apply Lemma~\ref{lem-Riesz-Thorin} with $\mathcal{T} = \mathcal{A}^{-1}\vert_{X}$ compact on $X$ to obtain the conclusion of Part~(b). {\rm Part~(c)}: Assume that $\lambda\in \mathbb{C}$ is not an eigenvalue of $\mathcal{A}$. With regard to Part~(a) we may restrict ourselves to the case $\lambda\not\in (-\infty, \Lambda)$. Hence, by the Riesz\--Schauder theory applied to $\mathcal{A}^{-1}$, which is compact on $L^2(\mathbb{R}^N)$, $\lambda$ is in the resolvent set of $\mathcal{A}$ and the resolvent $\mathcal{K} = (\mathcal{A} - \lambda I)^{-1}$ is compact on $L^2(\mathbb{R}^N)$. We refer to {Edwards} \cite[Theorem 9.10.2, p.~679]{Edwards} or {Yosida} \cite[Chapt.~X, Theorem 5.1, p.~283]{Yosida} for the Riesz\--Schauder theory. Consequently, the following identities hold on $L^2(\mathbb{R}^N)$: % \begin{equation} \mathcal{K} \left( \lambda^{-1} I - \mathcal{A}^{-1} \right) = \left( \lambda^{-1} I - \mathcal{A}^{-1} \right) \mathcal{K} = \lambda^{-1} \mathcal{A}^{-1}\, . \label{res:(A-lam)^-1} \end{equation} % In particular, $\lambda^{-1}$ cannot be an eigenvalue of $\mathcal{A}^{-1}$. So $\lambda^{-1}$ is not an eigenvalue of $\mathcal{A}^{-1}\vert_{X}$ either. The restriction $\mathcal{A}^{-1}\vert_{X}$ being compact on $X$, by Part~(a), we may apply Lemma~\ref{lem-Riesz-Thorin} with $\mathcal{T} = \mathcal{A}^{-1}\vert_{X}$ again to conclude that the restriction $\lambda^{-1} I - \mathcal{A}^{-1}\vert_{X}$ of $\lambda^{-1} I - \mathcal{A}^{-1}$ to $X$ has a bounded inverse, say, % \begin{math} \mathcal{L} = \left( \lambda^{-1} I - \mathcal{A}^{-1}\vert_{X} \right)^{-1}\, . \end{math} % Hence, from \eqref{res:(A-lam)^-1} we deduce % \begin{math} \mathcal{K}\vert_{X} = \lambda^{-1} \mathcal{L} \left( \mathcal{A}^{-1}\vert_{X} \right) \end{math} % which shows that also $\mathcal{K}\vert_{X}$ is compact on $X$ as claimed. The proof of {\rm Theorem~\ref{thm-X-compact}} is now complete. \end{proof} \subsection{Positivity for a nonradial potential $q(x)$} \label{ss:Positive-q(x)} \begin{proof}[Proof of Theorem~\ref{thm-Positive-q(x)}] Let $-\infty < \lambda < \Lambda_{q}$ and $u = (\mathcal{A}_{q} - \lambda I)^{-1}\vert_{X_{q}^\odot} f$. Since $0\leq f\in X_{q}^\odot$, we may apply the weak maximum principle % (as in the proof of Proposition~\ref{prop-Positive-Q(r)}) to get $0\leq u\in X_{q}^\odot$. Hence, it suffices to prove our theorem for $g = \min\{ f,\, \varphi_{q}\}$ in place of $f$, that is, for $0\leq f\leq \varphi_{q}$ a.e.\ and $f\not\equiv 0$ in $\mathbb{R}^N$. This forces also $0\leq u\leq (\Lambda_{q} - \lambda)^{-1} \varphi_{q}$ a.e.\ and $u\not\equiv 0$ in $\mathbb{R}^N$, by the weak maximum principle again. Similarly as in the proof of Theorem~\ref{thm-X-compact}, Part~(a) above, let us rewrite the equation $\mathcal{A}_{q} u = \lambda u + f$ for $u\in X_{q}$, with $f\in X_{q}$, $X_{q} = X_{Q_1} = X_{Q_2}$, as % \begin{equation*} - \Delta u + Q_2(|x|) u = \lambda u + g(x) \quad\mbox{in }\, X_{Q_2}^\odot , %\label{Q_2(r):Schroed_lam} \end{equation*} % where % \begin{equation*} g(x) = \left[ Q_2(|x|) - q(x) \right] u(x) + f(x) \geq f(x) ,\quad x\in \mathbb{R}^N , \end{equation*} % by condition \eqref{growth:q(x)} and $u\geq 0$ a.e.\ in $\mathbb{R}^N$. Again, we combine Corollary~\ref{cor-X-compact} with $$ \int_{\mathbb{R}^N} Q_2\, \varphi_{Q_2}^2 \,{\rm d}x < \infty $$ to get $\int_{\mathbb{R}^N} Q_2\, u\, \varphi_{Q_2} \,{\rm d}x < \infty$, that is, % \begin{math} Q_2\, u\in X_{Q_2}^\odot =\break L^1(\mathbb{R}^N; \varphi_{Q_2} \,{\rm d}x) . \end{math} % Consequently, also $(Q_2 - q) u\in X_{Q_2}^\odot$ which guarantees $g\in X_{Q_2}^\odot$. We apply Proposition~\ref{prop-Positive-Q(r)} with $Q=Q_2$ and $\lambda < \Lambda_{q}\leq \Lambda_Q = \Lambda_{Q_2}$ to conclude that $\inf_{\mathbb{R}^N} (u / \varphi_{Q_2}) > 0$ or, equivalently, $u\geq c\varphi_{q}$ a.e.\ in $\mathbb{R}^N$, with some constant $c\equiv c(f) > 0$. \end{proof} \section{Appendix} \label{s:Appendix} \subsection{An example of a monotone radial potential} \label{ss:Examples} Here we give an example of a radially symmetric potential $q(x) = q(r)$ which belongs to class \eqref{class:Q(r)}, but does {\it not } belong to the analogue of this class defined in {Alziary}, {Fleckinger}, and {Tak\'a\v{c}} \cite[eqs.\ (13) and (14)]{AFT-3}. More precisely, it fails to satisfy condition \eqref{e:int-deriv} for {\it any } $\gamma > 0$ (\cite[eq.~(14)]{AFT-3}). This example illustrates how ``large'' class \eqref{class:Q(r)} actually is. For $r\in \mathbb{R}_+$ we set either $q'(r) = 0$ or else $q'(r) = 2\, q(r)^{3/2}$, which yields a ``very fast'' growth of $q(r)$ on a sequence of pairwise disjoint, nonempty intervals $(n - \varrho_n, n + \varrho_n)$; $n=1,2,3,\dots$, of total length $2\sum_{n=1}^\infty \varrho_n = 1$, where $\varrho_n\to 0$ sufficiently fast as $n\to \infty$, say, $\varrho_n = \mathcal{O}(1/n^3)$. \begin{example}[{\cite[Example 3.6]{AFT-3}}] \label{exam-q(r):mono} \rm We define $q: \mathbb{R}_+\to (0,\infty)$ by $q(r) = \theta(r)^{-2}$ for $r\in \mathbb{R}_+$, where $\theta: \mathbb{R}_+\to (0,1]$ is a monotone decreasing, piecewise linear, continuous function defined as follows: Let $\{ \varrho_n\}_{n=1}^\infty \subset (0,1/2)$ be a sequence of numbers satisfying % \begin{equation} \sum_{n=1}^\infty \varrho_n = {1}/{2} \,. \label{sum:rho_n=1/2} \end{equation} % Given $r\geq 0$, we set $\theta(0) = 1$ and % \begin{equation*} %\label{e:deriv=-1/0} \frac{\mathrm{d}\theta}{\mathrm{d}r}(r) = \begin{cases} -1 & \mbox{if $|r-n| < \varrho_n$ for some $n\in \mathbb{N}$;} \\ 0 & \mbox{otherwise,} \\ \end{cases} \end{equation*} % where $\mathbb{N} = \{ 1,2,3,\dots\}$. Setting $R_0 = 1$ and abbreviating % \[ R_n = 1 - 2\sum_{k=1}^n \varrho_k > 0 \quad\mbox{for } n=1,2,\dots , %\label{sum:R_n} \] % we compute for $r\geq 0$: % \begin{equation*} \theta(r) = \begin{cases} 1 & \mbox{if } 0\leq r\leq 1 - \varrho_1 ; \\ R_{n-1} - ((r-n) + \varrho_n) & \mbox{if $|r-n| < \varrho_n$ for some $n\in \mathbb{N}$;} \\ R_n &\mbox{if } \varrho_n\leq r-n\leq 1 - \varrho_{n+1} \mbox{ for some $n\in \mathbb{N}$.} \end{cases} \end{equation*} % Clearly, $\theta: \mathbb{R}_+\to (0,1]$ is monotone decreasing, piecewise linear, and continuous. It satisfies $\theta(r)\to 0$ as $r\to 0+$, by \eqref{sum:rho_n=1/2}. Next, we compute % \begin{equation*} %\label{sum:int-Q^-1/2} \begin{aligned} \int_{1 - \varrho_1}^{\infty} \theta(r) \,\mathrm{d}r & = \sum_{n=1}^\infty (R_{n-1} - \varrho_n) \cdot 2\varrho_n + \sum_{n=1}^\infty R_n (1 - \varrho_{n+1} - \varrho_n) \\ & < 1 + 2\sum_{n=1}^\infty R_n \,. \end{aligned} \end{equation*} % Furthermore, for any $\gamma > 0$ we get % \begin{equation*} %\label{sum:int-deriv} \begin{aligned} & \int_{0}^{\infty} \left| \frac{{\rm d}}{{\rm d}r} \left( q(r)^{-1/2} \right) \right|^{\gamma} q(r)^{1/2} \,{\rm d}r \\ & = \sum_{n=1}^\infty \int_{|r-n| < \varrho_n} \left[ R_{n-1} - ((r-n) + \varrho_n) \right]^{-1} \,\mathrm{d}r > 2\sum_{n=1}^\infty \varrho_n\, R_{n-1}^{-1} \,. \end{aligned} \end{equation*} % We will have an example of a potential $q(r)$ with the desired properties as soon as we find a sequence $\{ \varrho_n\}_{n=1}^\infty \subset (0,1/2)$ that satisfies the following conditions: % \begin{equation} \sum_{n=1}^\infty \varrho_n = {1}/{2} \,,\quad %\label{sum:rho_n=1/2} \sum_{n=1}^\infty R_n < \infty \,,\quad\mbox{and }\quad %\label{sum:R_n 0$, the function $S(t) = \gamma\cdot \exp( \beta t^{\alpha} ) > 0$ obeys inequality \eqref{e:exp-power-gr} with $C = \alpha_1 ( 1 + \log^+ \gamma^{-1} )$ where $\alpha_1 = \max\{ \alpha, 1\}$. % \end{lemma} \begin{proof} From \eqref{e:exp-power-gr} we deduce % \begin{equation*} \begin{split} \int_{r_0}^s \exp\Big( - \int_r^s S(t) \,\mathrm{d}t \Big) \,\mathrm{d}r &\leq \int_{r_0}^s \exp\Big( - \,\frac{(s-r) S(s)}{ C\, (1 + \log^+ S(s)) }\Big) \,\mathrm{d}r \\ & = \frac{ C\, (1 + \log^+ S(s)) }{S(s)} \Big[ 1 - \exp\Big( - \,\frac{(s-r) S(s)}{ C\, (1 + \log^+ S(s)) } \Big) \Big] \\ & \leq \frac{ C\, (1 + \log^+ S(s)) }{S(s)} \end{split} \end{equation*} % for all $s\geq r_0$. Consequently, \eqref{e:int-diff} follows from % \begin{equation*} \begin{aligned} & \int_{r_0}^\infty ( Q_2(s) - Q_1(s) ) \int_{r_0}^s \exp \Big( - \int_r^s [ Q_1(t)^{1/2} + Q_2(t)^{1/2} ] \,\mathrm{d}t \Big) \,\mathrm{d}r \,\mathrm{d}s \\ & \leq \int_{r_0}^\infty ( Q_2(s) - Q_1(s) ) \,\frac{ C\, (1 + \log^+ S(s)) }{S(s)}\, \,\mathrm{d}s \\ & = C\int_{r_0}^\infty \left( Q_2(s)^{1/2} - Q_1(s)^{1/2} \right) \left[ 1 + \log^+ ( Q_1(r)^{1/2} + Q_2(r)^{1/2} ) \right] \,\mathrm{d}r < \infty \,, \end{aligned} \end{equation*} % by condition \eqref{exp:int-diff}. We leave the example $S(t) = \gamma\cdot \exp( \beta t^{\alpha} )$ to the reader as an easy exercise. \end{proof} \subsection{Extensions of certain symmetric operators} \label{ss:Exten-oper} We present a few obvious, but necessary facts about extensions of bounded symmetric operators defined on $X$ to $L^2(\mathbb{R}^N)$ and $X^\odot$. The following lemma is an easy consequence of the Riesz\--Thorin interpolation theorem ({M.~Reed} and {B.~Simon} \cite[Sect.\ IX.4, Theorem IX.17, p.~27]{RSimon-II}). It is applied to the resolvent $\mathcal{K} = (\mathcal{A} - \lambda I)^{-1}$ on $L^2(\mathbb{R}^N)$, for $\lambda < \Lambda$, which is bounded on $X$ by inequality \eqref{e:max-princ_X}, and to similar operators as well. \begin{lemma}\label{lem-Riesz-Thorin} Let $q$, $\varphi$, $X$, and $X^\odot$ be as in {\rm Section~\ref{s:Main}}. Assume that $\mathcal{T}: X\to X$ is a bounded linear operator that satisfies the symmetry condition % \begin{equation} \int_{\mathbb{R}^N} (\mathcal{T} f)\, \bar{g} \,{\rm d}x = \int_{\mathbb{R}^N} f\, \overline{(\mathcal{T} g)} \,{\rm d}x \quad\mbox{for all } f,g\in X . \label{symm:T_X} \end{equation} % Then $\mathcal{T}$ possesses a unique extension $\mathcal{T}\vert_{X^\odot}$ to a bounded linear operator on $X^\odot$, $\mathcal{T}$ is the adjoint of $\mathcal{T}\vert_{X^\odot}$, and $\mathcal{T}\vert_{X^\odot}$ restricts to a bounded selfadjoint operator $\mathcal{T}\vert_{L^2(\mathbb{R}^N)}$ on $L^2(\mathbb{R}^N)$. Moreover, the operator norms of $\mathcal{T}\vert_{L^2(\mathbb{R}^N)}$, $\mathcal{T}\vert_{X^\odot}$, and $\mathcal{T}$, respectively, satisfy % \begin{equation} \| \mathcal{T}\vert_{L^2(\mathbb{R}^N)} \|_{ L^2(\mathbb{R}^N)\to L^2(\mathbb{R}^N) } \leq \| \mathcal{T}\vert_{X} \|_{X^\odot\to X^\odot} = \| \mathcal{T} \|_{X\to X} < \infty \,. \label{ineq-Riesz-Thorin} \end{equation} % The spectrum of $\mathcal{T}\vert_{L^2(\mathbb{R}^N)}$ is contained in the spectrum of $\mathcal{T}$. Finally, if $\mathcal{T}$ is compact, then so is $\mathcal{T}\vert_{L^2(\mathbb{R}^N)}$ and their spectra coincide; in particular, if $\mathcal{T}\vert_{L^2(\mathbb{R}^N)} v = \lambda v$ for some $\lambda\in \mathbb{C}\setminus \{ 0\}$ and $v\in L^2(\mathbb{R}^N)\setminus \{ 0\}$, then $\lambda\in \mathbb{R}$ and $v\in X$. % \end{lemma} This lemma is proved in {Alziary}, {Fleckinger}, and {Tak\'a\v{c}} \cite[Lemma 4.3]{AFT-3}. \subsection*{Acknowledgment} The authors would like to express their sincere thanks to {\it Professor Jacqueline Fleckinger-Pell\'e} (Universit\'e Toulouse~1, France) for her tremendous support of their joint work on this and a number of other papers about Schr\"odinger operators. Since 1993 both authors have enjoyed working with her and obtaining numerous interesting scientific results. During the pastime they have appreciated a lasting friendship with her, her readiness to help, and her everlasting optimism that she has always been willing to share with many others. This work of all three authors was supported in part by le Minist\`ere des Affaires \'Etrang\`eres (France) and the German Academic Exchange Service (DAAD, Germany) within the exchange program ``PROCOPE''. A part of this research was performed when P.~Tak\'a\v{c} was a visiting professor at CEREMATH -- UMR MIP, Universit\'e Toulouse~1, Toulouse, France. \begin{thebibliography}{99} \bibitem{AFT-1} B.~Alziary, J.~Fleckinger and P.~Tak\'a\v{c}; \emph{An extension of maximum and anti-maximum principles to a Schr\"odinger equation in $\mathbb{R}^2$}, J.~Differential Equations, {\bf 156} (1999), 122--152. \bibitem{AFT-2} B.~Alziary, J.~Fleckinger and P.~Tak\'a\v{c}; \emph{Positivity and negativity of solutions to a Schr\"o\-din\-ger equation in $\mathbb{R}^N$}, Positivity, {\bf 5}(4) (2001), 359--382. \bibitem{AFT-3} B.~Alziary, J.~Fleckinger and P.~Tak\'a\v{c}; \emph{Ground\--state positivity, negativity, and compactness for a Schr\"odinger operator in $\mathbb{R}^N$}, J.~Funct. Anal., {\bf 245} (2007), 213--248. {\it Online}: doi: 10.1016/j.jfa.2006.12.007. \bibitem{AlziaryTak} B.~Alziary and P.~Tak\'a\v{c}; \emph{A pointwise lower bound for positive solutions of a Schr\"odinger equation in $\mathbb{R}^N$}, J.~Differential Equations, {\bf 133}(2) (1997), 280--295. \bibitem{ClemPel} Ph.~Cl\'ement and L.~A. Peletier; \emph{An anti-maximum principle for second order elliptic operators}, J.~Differential Equations, {\bf 34} (1979), 218--229. \bibitem{Davies} E.~B. Davies; {\sl ``Heat Kernels and Spectral Theory''}, Cambridge Univ.\ Press, Cambridge, 1989. \bibitem{DavSimon} E.~B. Davies and B.~Simon; \emph{Ultra\-contractivity and the heat kernel for Schr\"odinger operators and Dirichlet Laplacians}, J.~Funct. Anal., {\bf 59} (1984), 335--395. \bibitem{Edwards} R.~E. Edwards; {\sl ``Functional Analysis: Theory and Applications''}, Holt, Rinehart and Winston, Inc., New York, 1965. \bibitem{EdmEvans} D.~E. Edmunds and W.~D. Evans; {\sl ``Spectral Theory and Differential Operators''}, Oxford University Press, Oxford, 1987. \bibitem{LCEvans} L.~C. Evans; {\sl ``Partial Differential Equations''}, Graduate Studies in Mathematics, Vol.~{\bf 19}, American Math.\ Society, Providence, R.~I., 1998. \bibitem{GilbargTrud} D.~Gilbarg and N.~S. Trudinger; {\sl ``Elliptic Partial Differential Equations of Second Order''}, Springer-Verlag, New York--Berlin--Heidelberg, 1977. \bibitem{Hartman} Ph.~Hartman; {\sl ``Ordinary Differential Equations''}, 2nd Ed., Birkh\"auser, Boston--Basel--Stuttgart, 1982. \bibitem{Hartman-Wint} Ph.~Hartman and A.~Wintner; \emph{Asymptotic integrations of linear differential equations}, American J. Math., {\bf 77}(1) (1955), 45--86. \bibitem{HoffSwetina} M.~Hoffmann-Ostenhof, T.~Hoffmann-Ostenhof, and J.~Swetina, \emph{Pointwise bounds on the asymptotics of spherically averaged $L^2$-solutions of one-body Schr\"odinger equations}, Ann. Inst. Henri Poincar\'e Anal. Physique th\'eorique, {\bf 42}(4) (1985), 341--361. \bibitem{Hoffmann} M.~Hoffmann-Ostenhof; \emph{On the asymptotic decay of $L^2$-solutions of one-body Schr{\"o}\-din\-ger equations in unbounded domains}, Proc. Roy. Soc. Edinburgh, {\bf 115A} (1990), 65--86. \bibitem{Kato} T.~Kato; {\sl ``Perturbation Theory for Linear Operators''}, in \emph{Grundlehren der ma\-the\-ma\-ti\-schen Wissenschaften}, Vol.~{\bf 132}. Springer\--Verlag, New York\--Berlin\--Hei\-del\-berg, 1980. \bibitem{ProtterWein} M.~H. Protter and H.~F. Weinberger; {\sl ``Maximum Principles in Differential Equations''}, Springer-Verlag, New York--Berlin--Heidelberg, 1984. \bibitem{RSimon-II} M.~Reed and B.~Simon; {\sl ``Methods of Modern Mathematical Physics, Vol.~II: Fourier Analysis, Self-Adjointness''}, Academic Press, Inc., Boston, 1975. \bibitem{RSimon-IV} M.~Reed and B.~Simon; {\sl ``Methods of Modern Mathematical Physics, Vol.~IV: Analysis of Operators''}, Academic Press, Inc., Boston, 1978. \bibitem{Sweers} G.~Sweers; \emph{Strong positivity in $C(\overline{\Omega})$ for elliptic systems}, Math. Z., {\bf 209} (1992), 251--271. \bibitem{Takac} P.~Tak\'a\v{c}; \emph{An abstract form of maximum and anti-maximum principles of Hopf's type}, J. Math. Anal. Appl., {\bf 201} (1996), 339--364. \bibitem{Titchmarsh} E.~C. Titchmarsh; {\sl ``Eigenfunction Expansions Associated with Second\--order Differential Equations''}, Part I, Oxford University Press, Oxford, England, 1962. \bibitem{Yosida} K.~Yosida; {\sl ``Functional Analysis''}, 6th Ed., in \emph{Grundlehren der ma\-the\-ma\-ti\-schen Wissenschaften}, Vol.~{\bf 123}. Springer\--Verlag, New York\--Berlin\--Hei\-del\-berg, 1980. \end{thebibliography} \end{document}